Next Article in Journal
Recent Progress in Catalytically Driven Advanced Oxidation Processes for Wastewater Treatment
Previous Article in Journal
Macroporous Resin-Based La-N Co-Doped TiO2 Composites for Efficient Removal of Environmental Pollutants in Water via Integrating Adsorption and Photocatalysis
Previous Article in Special Issue
CO2 Methanation over Ni-Based Catalysts: Investigation of Mixed Silica/MgO Support Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Calcination and Reduction Conditions of Ni-Al-LDH Catalysts for CO2 Methanation

by
Nailma Martins
and
Oscar W. Perez-Lopez
*
Laboratory of Catalytic Processes-PROCAT, Department of Chemical Engineering, Federal University of Rio Grande Do Sul (UFRGS), Ramiro Barcelos Street, 2777, Porto Alegre CEP 90035-007, Brazil
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(8), 760; https://doi.org/10.3390/catal15080760
Submission received: 30 May 2025 / Revised: 3 July 2025 / Accepted: 18 July 2025 / Published: 8 August 2025
(This article belongs to the Special Issue Catalysis and Technology for CO2 Capture, Conversion and Utilization)

Abstract

CO2 methanation offers a sustainable route to reduce greenhouse gas emissions by converting carbon dioxide into methane, a valuable renewable fuel. This exothermic reaction not only mitigates its environmental impact but also provides energy-efficient benefits, as the heat generated can be reused in industrial applications. In this study, CO2 methanation was carried out in a continuous flow reactor with a CO2:H2 molar ratio of 1:4 and a gas hourly space velocity (GHSV) of 12,000 h−1, using a Ni-Al-LDH catalyst with a molar ratio of 2.3. The research focused on how calcination and reduction conditions affect catalyst structure and activity. Characterization techniques such as BET, XRD, TPR, H2-TPD, and CO2-TPD revealed that these conditions significantly influence surface area, crystallinity, phase composition, and metal dispersion. A higher reduction temperature decreased the surface area and increased both the crystallite size and basicity. The findings highlight that thermal treatment play a crucial role in optimizing the catalytic properties of NiAl catalyst. The sample calcined at 600 °C showed greater activity at lower reaction temperatures, while the catalyst calcined at 400 °C performed better above 300 °C. Additionally, the evaluation of the effect of the reduction atmosphere during catalyst activation showed that H2 is a more effective reducing gas at lower reaction temperatures, whereas biogas showed a better performance at higher temperatures. Importantly, XRD results showed the catalysts maintained their structural integrity post-reaction, with no significant carbon deposition in the H2 atmosphere, confirming their potential for long-term application in CO2 methanation.

Graphical Abstract

1. Introduction

CO2 methanation is a promising solution to address environmental, technological, and economic challenges related to greenhouse gas emissions and the energy transition. By converting CO2 into methane, it enables carbon capture and reuse, mitigating climate change while producing a renewable energy source that can replace fossil fuels and balance the intermittency of solar and wind energy [1]. Methanation also supports power-to-gas technologies for large-scale energy storage, provides feedstock for the chemical industry, and improves the utilization of green hydrogen produced via water electrolysis [2,3,4]. Its energy efficiency, economic viability, and alignment with global decarbonization goals make it a key technology for the circular economy and sustainable development.
The CO2 conversion reaction is a catalytic process that involves the conversion of CO2 and hydrogen (H2) into CH4 and water (H2O) over a suitable catalyst under low temperatures and pressures [5]. The chemical reaction can be represented as CO2 + 4H2 → CH4 + 2H2O and is known as the Sabatier reaction whose limitations are slow kinetics because the C=O bond energy is high (750 kJ/mol), andCO2 dissociation is very difficult [5,6,7]. This exothermic reaction produces renewable methane, which can be integrated into the existing natural gas infrastructure or serve as a feedstock for various chemical processes. Furthermore, the generation of low amounts of by products and high conversion rates when using a catalyst, makes the methanation process a promising large-scale method [2,8].
The catalyst is a crucial component of methanation, which explains why numerous researchers have devoted their efforts to synthesizing and investigating various materials for the Sabatier reaction [9,10,11,12]. Noble metal catalysts consistently exhibit exceptional performance in CO2 methanation [13]. Nevertheless, as economic considerations necessitate the exploration of cost-effective and readily available alternatives, Ni-based catalysts are employed due to their low cost and good performance [8,9,10,11]. The structure of Ni-based catalysts is a crucial factor influencing the performance of CO2 methanation. Given the structure-sensitive nature of the reaction, Ni particle size significantly impacts catalyst effectiveness. However, nickel catalysts demonstrate lower activity at low temperatures relative to Rh and Ru-based catalysts. Additionally, the presence of CO as a byproduct can lead to the formation of nickel carbonyl species and carbon deposition on the catalyst surface, which can result in deactivation [9,14,15].
Various methodologies have been utilized in the development of nickel-based catalysts, with a focus on controlling basicity, appropriate nickel reducibility, and particle size, all of which are crucial for adsorption capacity and the CO2 redox reaction [11,12]. The currently developed synthesis route involves the formation of layered double hydroxide (LDH), which contain mixtures of divalent/trivalent metals comprising polycations and feature a layered structure. LDHs have several interesting properties, such as their basic character, ion exchange capacity and memory effect by which their structure can be reconstructed via moisture after calcination [9,12,14].
In a previous work some LDH catalysts were prepared by our group via co-precipitation [9,16]. Martins et al. 2025 [9] show that modifying the amount of aluminum and nickel in the structure of LDHs can increase the surface area and dispersion of the catalysts, resulting in a greater conversion of CO2 to methane. Ni-Al-LDH catalysts were evaluated in a broad Ni/Al range (from 1.51 to 3.06) for CO2 methanation, where the Ni70Al30 catalyst showed the best performance and stability. According to Perez-Lopez (2006) [16], the reduction temperature significantly influences the activity and selectivity of the catalysts for the CO2 reforming of methane. In contrast, the catalytic properties were practically independent of the calcination temperature, which primarily affects the surface area, which decreases as the calcination temperature increases.
Considering that calcination has a significant impact on the structure and composition of LDHs and greatly improves their adsorption performance, in this study, the effect of calcination temperature and reduction conditions (temperature and atmosphere) on the physicochemical and catalytic properties of the Ni70Al30 catalyst for CO2 methanation was evaluated.

2. Results

2.1. Catalytic Characterization

The elemental analysis was performed as previously outlined [9], where the Ni/Al molar ratio obtained by ICP-OES was 2.32, very close to the nominal value (2.3), with a deviation of less than 1%. Furthermore, the weight % of the experimental results in XRF were 81.4 wt% and 18.6 wt% for Ni and Al, respectively, with deviation below 3%. These results revealed that the real values were very close to the nominal ones, demonstrating the reproducibility of the coprecipitation method.
Figure 1 presents the N2 adsorption–desorption isotherms, which are used to evaluate the surface properties of the samples. According to the IUPAC classification [17], the isotherms for all samples calcined at different temperatures are categorized as Type IV, with an H3-type hysteresis loop. This formation is typical of hysteresis indicating capillary condensation, which is characteristic of mesoporous materials with pore sizes ranging between 2 and 50 nm.
The experimental data related to nitrogen adsorption and desorption, such as the BET surface area, the average pore size, and the average pore volume are presented in Table 1. The BET analysis suggests that an increase in calcination temperature results in a decrease in the surface area. However, this change in temperature does not appear to affect the pore volume. This observation may be crucial to understand the optimization of catalysts for specific applications, in which the surface area is a determining factor in catalytic activity. Pinthong et al. (2019) [18] demonstrated in their studies with Mg-Al Layered Double Hydroxides that the BET surface area gradually decreased as the calcination temperature increased. The average pore size diameters of all samples ranged between 12 and 20 nm. Additionally, as the calcination temperature increased, the pore volumes decreased, which in turn resulted in an increase in the average pore size diameter.
The calcination of nickel hydrotalcites (Ni-LDHs) has a significant impact on their structural and catalytic properties. Calcination leads to the decomposition of the lamellar structure of hydrotalcite, resulting in the formation of mixed oxides of nickel and aluminum (Ni-Al-O) and this effect can be observed in the XRD results presented in Figure 2. The calcined samples present reflections at 2θ angles = 37.3°, 43.2°, and 62.8°, respectively, and can be identified as the (111), (200) and (220) crystal planes in the face-centered cubic crystalline structure of NiO [9]. However, Ni-Al-O spinels also exhibit a similar diffraction pattern to that of NiO, and the peak at 37.5° may be associated with the phases of Ni-Al-O spinels (NiAl2O4–normal and Ni2AlO4–inverse) [9,19]. There are no reflections that could be associated with other crystalline compounds such as simple metal hydroxides.
The crystallite sizes of the calcined materials were determined from the reflection of the (200) plane of NiO. The sample calcined at 800 °C exhibited the largest crystallite size (4.4 nm), while the sample calcined at 400 °C presented the smallest crystallite size (3.0 nm). Calcination can decrease the dispersion of nickel particles, leading to larger crystallite sizes [20,21]. Similarly, smaller crystallite sizes generally result in a larger surface area. This effect was confirmed by a BET analysis (Table 1), which showed that samples with larger crystallite sizes had smaller surface areas.
The three samples obtained by calcination at different temperatures and reduced at 600 °C presented the diffraction profile shown in Figure 2b. All samples presented a small peak at 36.6°, mainly in the sample calcined at a low temperature (400 °C), indicating that the mixed oxides were not fully reduced. The presence of metallic nickel was evident in all three samples after the reduction process, with peaks at angles of 44.4°, 51.5°, and 76.3°, corresponding to the (111), (200), and (220) crystal planes of Ni0, respectively. Clearly, the crystallinity of the samples increases with the calcination temperature. The increase in crystallite size after reduction can be explained by the exposure of the metallic area and grain growth. The formation of metallic nickel at lower temperatures, particularly in the sample calcined at 400 °C, can be attributed to the presence of weakly bound NiO species, which are more easily reducible.
The temperature-programmed reduction (TPR) method was employed to investigate the reducibility of the catalyst, which is strongly influenced by the interaction strength between the metal and the structure (Figure 3). The calcination temperature significantly affects the reduction behavior of catalysts [22,23]. The TPR profile for the sample calcined at 600 °C for 6 h exhibited the thinnest peak, indicating a more uniform distribution of particle sizes. In contrast, the samples calcined at 400 °C and 800 °C showed broader peaks, suggesting less uniformity and more dispersed particles.
The wide range of reduction temperatures indicates the presence of different nickel oxide species [9,23]. The profiles were refined through deconvolution, and the temperature ranges were used to determine the reduced species. Bulk NiO, being an easily reducible phase, is the first to undergo reduction at lower temperatures between 250 °C and 400 °C. Smaller particles with larger surface areas tend to reduce more rapidly due to the increased contact area with the reducing gas. This effect is evident in the sample calcined at 400 °C, which exhibited a smaller crystallite size and greater surface area, and this phase represents approximately 50% for this sample. The sample calcined at 600 °C did not show a reduction peak at temperatures below 500 °C, whereas the sample calcined at 800 °C contained approximately 20% of this phase.
The reduction peaks between 530 °C and 720 °C indicate a reduction in mixed Ni-Al oxides. These peaks are attributed to the reduction in the inverse spinel (Ni2AlO4) and normal spinel (NiAl2O4), corresponding to the third and four peaks, respectively. This result indicates that calcination at different temperatures significantly influenced the nickel species formed during the reduction, potentially leading to variations in catalytic performance.
The sample calcined at 600 °C exhibited higher reduction temperatures for all three reduction peaks, indicating more stable nickel species, compared to the samples calcined at 400 °C or 800 °C. Furthermore, this sample showed a larger metallic area, along with the largest dispersion.
Sikander et al. 2019 [24] evaluated the different reducible species in samples with varying nickel contents and found that NiO is reduced at temperatures below 500 °C, while NiO interacting with the support requires a higher temperature for a complete reduction. The reduction of the Ni2+ species depends on the demographic position of the Ni species, whether in a non-stoichiometric spinel or mixed-oxide matrix, which are reduced at lower temperatures compared to nickel in the spinel-phase (Ni,Mg)Al2O4.
The H2-TPD profiles of the sample calcined at 400 °C for 8 h (Figure 4a) revealed three distinct regions of hydrogen desorption, indicating different metal-hydrogen interactions. Desorption at 191 °C is associated with hydrogen interacting with nickel reduced from NiO. Between 300 and 330 °C, hydrogen is adsorbed at active sites of NiAl spinels [25]. Above 600 °C, a broad peak indicates a strong hydrogen interaction with Ni-Al species. These observations are crucial for understanding catalytic behavior and optimizing catalyst performance in various reactions. The samples calcined at 600 °C and 800 °C did not exhibit a reduction peak at low temperatures in the TPR analysis, indicating there is no presence of bulk NiO in these materials. In the H2 chemisorption analyses, these samples displayed only two desorption regions, which are characteristic of hydrogen’s interaction with nickel reduced from NiAl spinels. This is evidenced by the difference in the H2 desorption profile compared to the sample calcined at 400 °C. Furthermore, the size of the crystallites can influence the interaction between the metal and the support. Larger crystallites may lead to less metal dispersion, affecting the strength of the hydrogen interaction with the metal and, consequently, the observed desorption temperature [25].
Based on the integration results of the obtained profile, the metallic surface area (SNi) and dispersion (γNi) were estimated according to the previously described methodology [9,25], and presented in Table 2. As observed, the total dispersion of metallic Ni decreases in the sample calcined at 800 °C, along with the metallic area, as a result of the sintering of dispersed nickel. These results show that the sample calcined at 800 °C has the lowest metallic area and least total dispersion, likely due to the sintering effect at higher temperatures. The sample calcined at 600 °C showed the highest metallic area and greatest dispersion, indicating a better distribution of nickel species.
As has already been clearly demonstrated in this work, the calcination process had a great influence on the Ni phases of the catalyst, presenting different reduction properties. The CO2-TPD profiles of the samples after the reduction in a 10% H2/N2 atmosphere at 600 °C are shown in Figure 5. The results for the three samples showed four desorption peaks at similar temperatures, the deconvolution of these profiles were compiled in Table 3. As reported in the literature, the basic sites on the catalyst surface can be classified into three types: weak basic sites located at about 160 °C, medium basic sites from 200 °C to 400 °C, and strong basic sites above 400 °C [26,27]. The basicity of mixed oxides derived from hydrotalcite is strongly dependent on the layer compositions, the presence of promoters, and the type of anions present between the interlayer spaces [28,29].
The presence of basic sites favors the adsorption of CO2, facilitating its activation and subsequent surface reactions. Furthermore, control of basic sites is essential to favor the oxidation of carbon deposits that may form. In our previous study, we assessed that the molar ratio of nickel significantly impacted the basicity of the catalyst [9]. In the current work, all samples contained an identical amount of nickel, resulting in a similar basicity across the three samples. The sample calcined at 600 °C exhibited lower basicity, likely due to a more uniform and homogeneous arrangement of nickel oxides, as suggested by the TPR analysis. The first peak at 185–190 °C, observed in all three samples, corresponds to the desorption of CO2 from weak Brønsted groups (OH) [29]. The intermediate peak at 262–277 °C is due to bidentate carbonates formed in metal–oxygen pairs [26]. Desorption peaks above 430 °C are linked to CO2 bound to low-coordination oxygen anions, indicating strong basic sites [7,26]. An additional peak above 600 °C is attributed to the completion of the dehydroxylation process and the decomposition of carbonate anions formed during CO2 adsorption [30].
The distribution of basic sites plays a fundamental role in the adsorption and activation of CO2 during methanation. In this study, CO2-TPD analysis revealed the presence of weak, medium, strong, and very strong basic sites, each contributing differently to the catalytic process [7]. Weak basic sites, typically associated with surface hydroxyl groups, are responsible for reversible CO2 adsorption and are essential for initiating the reaction pathway. Medium-strength sites, often related to bidentate carbonate formation on metal–oxygen pairs, are believed to stabilize key intermediates and enhance the reaction kinetics [31]. Strong and very strong basic sites, attributed to low-coordination oxygen anions, can promote deeper CO2 activation, but may also lead to overly strong adsorption, potentially hindering desorption and reducing turnover frequency [25].
In addition, some authors demonstrated that the weak and medium basic sites are the key to enhancing the methanation activity catalysts [9,19,28,29,30,31,32]. Ren et al. (2024) [29] synthesized a series of Mg-Al hydrotalcites, which were calcined at various temperatures to prepare supports for Ni-based catalysts. The calcination temperature affected the distribution of basic sites, with higher amounts of weak and medium basic sites increasing CO2 adsorption. This improved the catalyst activity and stability in CO2 methanation.

2.2. Catalytic Tests

The CO2 hydrogenation activity of NiAl catalysts calcined at different temperatures was evaluated at reaction temperatures between 200 and 400 °C, as shown in Figure 6a. The influence of the calcination temperature on the catalytic performance is evident. The activity of the NiAl-600 catalyst is higher than that of NiAl-400 and NiAl-800 at lower reaction temperatures. However, above 300 °C, the CO2 conversion of NiAl-600 was similar to NiAl-800, whereas NiAl-400 showed greater conversion, indicating that NiAl-400 requires higher temperatures to be fully activated. Despite having the same Ni/Al molar ratio, the catalytic activity profiles agreed with the characterization results, which showed different surface areas, crystallite sizes, and reduction temperatures. The TPD-H2 and TPR results, along with the crystallite size measurements, reveal significant insights into the behavior of NiAl catalysts calcined at different temperatures. The sample calcined at 400 °C for 8 h exhibits the highest total H2 consumption and metallic area, indicating a well-dispersed nickel phase with smaller crystallites, which demonstrates that the greatest CO2 conversions were achieved at reaction temperatures above 300 °C. As the calcination temperature increases to 600 °C and 800 °C, the crystallite size grows, leading to a decrease in metallic area and total dispersion. This sintering effect at higher temperatures results in fewer active sites for hydrogen adsorption, as evidenced by the lower total H2 consumption in the sample calcined at 600 °C. The sample calcined at 800 °C shows a significant reduction in dispersion and metallic area, correlating with the larger crystallite size and the reduced interaction strength between the hydrogen and nickel. These findings underscore the importance of optimizing calcination conditions to maintain high levels of dispersion and catalytic activity.
There is no linear correlation between the activity (CO2 conversion) and the specific surface area (SBET), especially for reaction temperatures below 300 °C, as previously reported [16]. Above this temperature, the sample calcined at 400 °C showed greater activity, which may be related to its high SBET. However, the catalyst activity must be related to the crystallite size, and in this case, the catalyst calcined at 400 °C presented the smallest crystallite size, which would explain the greater activity of this catalyst for reaction temperatures above 300 °C.
The NiAl-600 catalyst, which exhibited a balanced distribution of weak and medium basic sites, showed superior activity at lower temperatures, suggesting that these sites are more effective in promoting CO2 conversion under mild conditions. In contrast, the NiAl-800 catalyst, with a higher proportion of strong basic sites, demonstrated lower activity, likely due to stronger CO2 binding and the limited surface mobility of intermediates. These findings reinforce the importance of tailoring the basic site distribution to optimize the catalytic performance of LDH-derived materials.
Furthermore, a CH4 selectivity of 100% was achieved with all catalysts over the entire temperature range.
After the CO2 methanation reaction, the catalysts were characterized by XRD (Figure 6b) and the results indicate that there is no evidence of carbon deposition, whereas the crystalline structure was preserved and no new phase was detected in the NiAl-400, NiAl-600 and NiAl-800. The Ni reflections indicate that there was no sintering according to the crystallite sizes calculated by the Scherrer equation, determined from reflection of Ni0 at 44.5°. The crystallite size of the spent samples did not show any significant difference in relation to the samples reduced before the reaction (fresh samples). This observation indicates the absence of sintering, thereby ensuring the stability of the catalyst structure.
Temperature-programmed oxidation (TPO) was performed on the spent catalysts, as shown in Figure 7. This analysis aimed to identify and quantify carbonaceous species deposited on the catalyst surface, providing valuable insights into their stability and efficiency. The initial weight loss observed below 160 °C was attributed to the thermal desorption of water and CO2 adsorbed during the reaction, with water being a primary product of the methanation process. Between 160 °C and 400 °C, a weight gain was recorded for all samples, corresponding to the reoxidation of metallic Ni0 crystallites that remained after the catalytic tests [16]. Differential thermal analysis (DTA) revealed slight variations in the peak reoxidation temperature, with the NiAl-800 catalyst showing the lowest degree of reoxidation. Above 400 °C, weight loss was associated with the oxidation of carbonaceous deposits.
The NiAl-400 catalyst exhibited the highest carbon deposition (2.31%), which correlates with its higher CO2 conversion at elevated temperatures. In comparison, NiAl-600 and NiAl-800 showed lower mass losses of 1.58% and 1.76%, respectively. All values remained below 3%, in agreement with previous findings, confirming the favorable basic properties of the catalyst [9].

2.3. Effect of Reduction Conditions on Catalysts

The reduction step of the catalyst plays a crucial role in the catalytic activity for CO2 methanation. This process helps to better dispersion of nickel particles, thereby increasing the active surface area available for the reaction [33]. Furthermore, as reported by various studies, the catalyst reduction can enhance resistance to coke formation [9,18,33,34]. However, it is important to note that reduction/activation at inappropriate temperatures may lead to the sintering of nickel particles, which in turn decreases the active surface area and, consequently, the efficiency of the catalyst [35].
In this research, the influence of the reduction temperature on the NiAl catalyst was evaluated by reducing at 400, 500, and 600 °C under hydrogen flow. The catalytic activity results indicate that the H2 flow was more effective in the reduction process at low reaction temperatures compared to other gas mixtures (containing CH4 and/or CO2). Furthermore, the catalysts reduced at 500 and 600 °C showed equivalent results, while the catalyst reduced with biogas at 600 °C exhibited a better performance at higher reaction temperatures. The samples reduced with CH4 and biogas showed a significant amount of coke on the catalyst surface (Figure 8b). Apparently, the carbon deposited on the catalyst surface during the reduction with CO2 and CH4 led to the catalyst’s deactivation. However, when reduced with biogas, a carbon degassing process occurs at high temperatures, enhancing its catalytic activity.
After the CO2 methanation reaction, the catalysts were characterized by XRD (Figure 8c). The results indicate that their reduced crystal structure was maintained, with peaks identified with the symbol * corresponding to NiO. Additionally, the signal at 25.9 °C evidences carbon deposition in the sample reduced with CH4. This reduction step resulted in a significant increase in crystallite size, as shown in Table 4, indicating sintering and coke formation on the catalyst surface. For the other reducing gases, the diffraction reflections for Ni0 (111) showed no sintering, as determined by the crystallite sizes calculated using the Scherrer equation, thereby ensuring the stability of the catalyst structure.

3. Materials and Methods

3.1. Catalyst Preparation

Ni70Al30-LDH catalyst was synthesized according to previously reported procedures [9,16,19]. In summary, co-precipitation was performed at 50 °C and pH 8.0 ± 0.1 continuously in a jacketed reactor using a nitrate solution (1 M) of Ni and Al, along with an alkaline solution (2 M) of Na2CO3 and NaOH (50/50 wt%) from Synth, Brazil both added dropwise. Nickel nitrate (Ni(NO3)2·6H2O, 97% P.A.) and aluminum nitrate (Al(NO3)3·9H2O, 98% P.A.) from Vetec, Brazil were used. After precipitation, the resultant solution was aged for 1 h at 60 °C under agitation, then washed with deionized water and filtered until the conductivity was below 50 µS. The material was dried overnight at 80 °C and sieved (32–42 mesh), yielding the as-prepared LDH. Calcination in synthetic air was carried out at 400 °C, 600 °C and 800 °C for 8 h, 6 h and 2 h, respectively, to form mixed oxides. For evaluation of reduction temperature, sample calcined at 600 °C was submitted at reduction temperatures of 400, 500 and 600 °C for 4 h, 2 h and 1h, respectively. In addition, different gas mixtures were used during the activation/reduction: H2, CH4, CO2 and CH4/CO2.

3.2. Catalyst Characterization

The physicochemical properties of the calcined samples were determined using nitrogen adsorption–desorption isotherms, and the tests were performed with a pore and surface analyzer (Quantachrome 4200e). Prior to analysis, the samples were degassed under vacuum for 3 h at 300 °C, and the measurements were carried out using liquid nitrogen at −196 °C. The specific surface area was obtained via the BET method, while the micro- and meso-pore volumes, as well as the pore diameter, were determined using the t-plot and Barrett–Joyner–Halenda (BJH) models [9,19,32].
The crystalline phases were characterized through X-ray diffractometry (XRD) using a Mini Flex diffractometer (Rigaku, 30 kV, 10 mA) equipped with a Cu-Kα radiation source (λ = 0.154 nm). This comprehensive analysis included as-synthesized materials (LDH), fresh samples (calcined and reduced), and spent samples (post-catalytic tests). The average crystallite size was estimated from the XRD patterns using the Scherrer equation (Equation (1)) [9,19,36].
d = k λ β c o s ( θ )
Where k = 0.9, λ is the Cu-Kα radiation wavelength (0.154 nm), β is the line broadening at half width of NiO (2 0 0) or Ni0 (1 1 1) facets, and θ is the corresponding angle.
Thermal analyses, including H2 temperature-programmed reduction (H2-TPR), CO2 temperature-programmed desorption (CO2-TPD), and H2 temperature-programmed desorption (H2-TPD) measurements, were conducted using a multipurpose instrument (SAMP3) equipped with a thermal conductivity detector (TCD). For H2-TPR measurements, approximately 100 mg of the samples were loaded into a U-tube quartz reactor and pretreated via N2 flow at 100 °C. The analyses were carried out from 100 to 850 °C (with a ramp rate of 10 °C/min) in a flow of 5% H2/N2 (30 mL/min) [9,19,32].
Prior to programmed temperature desorption characterization, the samples underwent reduction with 10% H2/N2 (100 mL/min) at 600 °C for 1 h. The basic properties of the samples were evaluated via CO2-TPD analyses. Reduced samples (100 mg) were heated to 100 °C, purged with He flow (30 min), then switched to CO2 (30 min, 30 mL/min) for the adsorption step, and purged with He (30 min). Finally, CO2 desorption was carried out from 100 to 800 °C (10 °C/min), with He as the carrier gas (30 mL/min) [9,19,37].
H2 chemisorption properties were evaluated via H2-TPD analyses. Around 100 mg calcined samples were reduced at 600 °C for 1 h with 10% H2/N2 (100 mL/min) flow, purged in N2 flow (30 min) at room temperature, switched to H2 flow (1 h, 20 mL/min) for the adsorption step, and purged with N2 (30 min, 30 mL/min). Lastly, H2 desorption was carried out from 50 °C to 800 °C (10 °C/min), with N2 as the carrier gas (30 mL/min). Surface metallic area and metal dispersion were calculated using Equations (2) and (3) [19,38]:
S N i 0 ( m 2 g 1 ) = Y × N A ×   F S A
γ N i 0 ( % ) = Y × F S W m M m × 10
with Y representing the H2 chemisorbed amount (mol/gcat), NA is the Avogadro number (6.023 × 1023 atoms/mol), A is the number of surface Ni atoms located at a unit area (1.54 × 1019 atoms/m2), FS is the stoichiometric factor (H2/Ni = 2), Wm is the Ni metal loading (gNi/gcat), and Mm is the Ni molar mass (58.69 gNi/mol) [9,19].

3.3. Catalyst Tests

Catalytic tests were conducted, as comprehensively detailed in previous studies [9,19,32]. Quartz wool was utilized to support the catalytic bed within a fixed-bed tubular quartz reactor, which was heated by an electric resistive furnace. Digital mass controllers (Sierra Instruments, Monterey, CA, USA) were employed to regulate gas flows. The outlet gases were analyzed online using a gas chromatograph (Varian Co., Model: Star 3600cx, Lexington, MA, USA) equipped with a thermal conductivity detector, a Porapak-T column, and nitrogen as the carrier gas.
The samples, typically 100 mg, were initially reduced in situ at 400, 500 and 600 °C for 4, 2 and 1 h, respectively, using a flow of 100 mL/min of reducing gas. During the potential reduction assessment stage, the gases used were hydrogen (10% in N2), CH4 (10% in N2), CO2 (10% in N2), and biogas (comprising 15% CH4 and 10% CO2 in N2).
All tests were conducted at atmospheric pressure with a total gas hourly space velocity (GHSV) of 60,000 mL/(gcat·h). Activity tests were carried out in stepwise mode with a feed composition of H2:CO2:N2 = 4:1:15 (v/v, 100 mL/min), varying the temperature from 200 to 400 °C in 50 °C increments, and with five gas chromatography (GC) analyses per temperature.
X C O 2 ( % ) = F C O 2 i n F C O 2 o u t F C O 2 i n × 100
S C H 4 ( % ) = F C H 4 o u t F C H 4 o u t + F C O o u t × 100
Y C H 4 % = X C O 2 . S C H 4 × 100
To investigate the presence of carbon deposits in spent samples, temperature-programmed oxidation (TPO) was conducted using a thermobalance (SDT Q600, TA Instruments, Newcastle, DE, USA). The samples (10 mg) were heated from ambient temperature to 800 °C at a rate of 10 °C/min in a synthetic air flow (100 mL/min) [9,16,19].

4. Conclusions

The N2 physisorption analysis revealed that increasing the calcination temperature decreases the surface area. The XRD analysis indicated that calcination leads to the formation of mixed nickel and aluminum oxides, with higher temperatures resulting in larger crystallite sizes and smaller surface areas.
The TPR profiles indicated that the calcination temperature significantly affects the reduction behavior of the catalysts, with the catalyst calcined at 600 °C showing the most uniform particle size distribution. CO2 desorption (CO2-TPD) highlighted the importance of basic sites in CO2 adsorption, facilitating its activation.
Calcination and reduction of Ni-Al-LDH catalysts has been shown to significantly impact the structural and catalytic properties in CO2 methanation.
The NiAl-600 catalyst showed higher activity at lower reaction temperatures, while the NiAl-400 catalyst performed better above 300 °C. All catalysts achieved 100% selectivity for CH4 across the entire temperature range. Thermogravimetry revealed that the NiAl-400 catalyst had the highest carbon deposition related to the highest CO2 conversion at high reaction temperatures.
Hydrogen proved to be a more effective reducing gas at lower reaction temperatures, while biogas showed a better performance at higher temperatures, indicating that the choice of reduction atmosphere can be tailored to the intended operating conditions.
These results highlight the importance of choosing the appropriate calcination temperature and reduction conditions to optimize the performance of Ni-Al-LDH catalysts in the methanation of CO2. A detailed understanding of the structural and catalytic properties of these materials is essential for the development of more efficient and sustainable processes for the conversion of CO2 to methane, contributing to climate change mitigation.

Author Contributions

Conceptualization, O.W.P.-L.; validation, N.M.; formal analysis, O.W.P.-L. and N.M.; investigation, N.M.; writing—original draft preparation, N.M.; writing—review and editing, O.W.P.-L.; supervision, O.W.P.-L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CAPES and CNPq Brazilian agencies.

Data Availability Statement

Data will be made available on request.

Acknowledgments

The authors are grateful for the financial support provided by CAPES (Brazilian Agency for Improvement of Graduate Personnel) and CNPq (National Council for Scientific and Technological Development).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ahmed, S.; Khan, M.K.; Kim, J. Revolutionary Advancements in Carbon Dioxide Valorization via Metal-Organic Framework-Based Strategies. Carbon Capture Sci. Technol. 2025, 15, 100405. [Google Scholar] [CrossRef]
  2. Medina, O.E.; Amell, A.A.; López, D.; Santamaría, A. Comprehensive Review of Nickel-Based Catalysts Advancements for CO2 Methanation. Renew. Sustain. Energy Rev. 2025, 207, 114926. [Google Scholar] [CrossRef]
  3. Tommasi, M.; Degerli, S.N.; Ramis, G.; Rossetti, I. Advancements in CO2 Methanation: A Comprehensive Review of Catalysis, Reactor Design and Process Optimization. Chem. Eng. Res. Des. 2024, 201, 457–482. [Google Scholar] [CrossRef]
  4. Mebrahtu, C.; Krebs, F.; Abate, S.; Perathoner, S.; Centi, G.; Palkovits, R. CO2 Methanation: Principles and Challenges. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 2019; pp. 85–103. [Google Scholar]
  5. Yarbaş, T.; Ayas, N. A Detailed Thermodynamic Analysis of CO2 Hydrogenation to Produce Methane at Low Pressure. Int. J. Hydrog. Energy 2024, 49, 1134–1144. [Google Scholar] [CrossRef]
  6. Chen, Y.; Zhang, Y.; Fan, G.; Song, L.; Jia, G.; Huang, H.; Ouyang, S.; Ye, J.; Li, Z.; Zou, Z. Cooperative Catalysis Coupling Photo-/Photothermal Effect to Drive Sabatier Reaction with Unprecedented Conversion and Selectivity. Joule 2021, 5, 3235–3251. [Google Scholar] [CrossRef]
  7. Ren, J.; Lou, H.; Xu, N.; Zeng, F.; Pei, G.; Wang, Z. Methanation of CO/CO2 for Power to Methane Process: Fundamentals, Status, and Perspectives. J. Energy Chem. 2023, 80, 182–206. [Google Scholar] [CrossRef]
  8. Sani, L.A.; Bai, H.; Xu, Z.; Fu, L.; Sun, Y.; Huang, X.; Gao, H.; Liu, X.; Bai, D.; Zhang, Z.; et al. Optimized Combustion Temperature in the Facile Synthesis of Ni/Al2O3 Catalyst for CO2 Methanation. J. CO2 Util. 2024, 80, 102678. [Google Scholar] [CrossRef]
  9. Martins, N.J.; Perez-Lopez, O.W. Tuning the Composition of Ni-Al-LDH Catalysts for Low-Temperature CO2 Methanation. Fuel 2025, 381. [Google Scholar] [CrossRef]
  10. Guo, J.; Yang, J.; Wang, Q.; Zhao, N.; Xiao, F. Effect of Ni-Ov-Ce Interface on Low Temperature CO2 Methanation. Fuel 2025, 381, 133568. [Google Scholar] [CrossRef]
  11. Han, H.; Zhang, S.; Song, S.; Zhang, W.; Liu, D.; Song, Z.; Wang, Q.; Ma, C.; Feng, S.; Duan, X. Construction of Bi2O3−x/NiAl-LDH S-Scheme Heterojunction for Boosting Photothermal-Assisted Photocatalytic CO2 Reduction. Appl. Surf. Sci. 2024, 662, 160122. [Google Scholar] [CrossRef]
  12. Wang, Z.; Zhang, T.; Ramirez Reina, T.; Huang, L.; Xie, W.; Musyoka, N.M.; Oboirien, B.; Wang, Q. Enhanced Low-Temperature CO2 Methanation over La-Promoted NiMgAl LDH Derived Catalyst: Fine-Tuning La Loading for an Optimal Performance. Fuel 2024, 366, 131383. [Google Scholar] [CrossRef]
  13. Sun, H.; Lv, J.; Wang, C.; Zhang, Y.; Sun, S.; Zhang, P.; Cheng, G.; Mei, D.; Wang, Y.; Yan, Z. Ru/CeO2 Catalysts with Enriched Oxygen Vacancies by Plasma Treatment for Efficient CO2 Methanation. Fuel 2025, 381, 133413. [Google Scholar] [CrossRef]
  14. Zhou, L.; Guo, X.; Hu, X.; Zhang, Y.; Cheng, J.; Guo, Q. CO2 Methanation Reaction over La-Modified NiAl Catalysts Derived from Hydrotalcite-like Precursors. Fuel 2024, 362, 130888. [Google Scholar] [CrossRef]
  15. Wasnik, C.G.; Nakamura, M.; Shimada, T.; Machida, H.; Norinaga, K. CO2 Methanation over Low-Loaded Ni-M, Ru-M (M = Co, Mn) Catalysts Supported on CeO2 and SiC. Carbon. Resour. Convers. 2024, 8, 100241. [Google Scholar] [CrossRef]
  16. Perez-Lopez, O.W.; Senger, A.; Marcilio, N.R.; Lansarin, M.A. Effect of Composition and Thermal Pretreatment on Properties of Ni–Mg–Al Catalysts for CO2 Reforming of Methane. Appl. Catal. A Gen. 2006, 303, 234–244. [Google Scholar] [CrossRef]
  17. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  18. Pinthong, P.; Praserthdam, P.; Jongsomjit, B. Effect of Calcination Temperature on Mg-Al Layered Double Hydroxides (LDH) as Promising Catalysts in Oxidative Dehydrogenation of Ethanol to Acetaldehyde. J. Oleo Sci. 2019, 68, 95–102. [Google Scholar] [CrossRef]
  19. Dias, Y.R.; Perez-Lopez, O.W. CO2 Methanation over Ni-Al LDH-Derived Catalyst with Variable Ni/Al Ratio. J. CO2 Util. 2023, 68, 102381. [Google Scholar] [CrossRef]
  20. Takehira, K. Highly Dispersed and Stable Supported Metal Catalysts Prepared by Solid Phase Crystallization Method. Catal. Surv. Asia 2002, 6, 19–32. [Google Scholar] [CrossRef]
  21. Takehira, K.; Kawabata, T.; Shishido, T.; Murakami, K.; Ohi, T.; Shoro, D.; Honda, M.; Takaki, K. Mechanism of Reconstitution of Hydrotalcite Leading to Eggshell-Type Ni Loading on MgAl Mixed Oxide. J. Catal. 2005, 231, 92–104. [Google Scholar] [CrossRef]
  22. Hidayat, T.; Rhamdhani, M.A.; Jak, E.; Hayes, P.C. The Kinetics of Reduction of Dense Synthetic Nickel Oxide in H2-N2 and H2-H2O Atmospheres. Metall. Mater. Trans. B 2009, 40, 1–16. [Google Scholar] [CrossRef]
  23. Yin, S.; Xu, C.; Yang, H.; Wu, C.; Wu, M.; Xu, J.; Zhu, H.; Qiu, J.; Xu, L.; Chen, M. Facilely Preparing Highly Dispersed Ni-Based CO2 Methanation Catalysts via Employing the Amino-Functionalized KCC-1 Support. Fuel 2024, 365, 131162. [Google Scholar] [CrossRef]
  24. Sikander, U.; Samsudin, M.F.; Sufian, S.; KuShaari, K.; Kait, C.F.; Naqvi, S.R.; Chen, W.-H. Tailored Hydrotalcite-Based Mg-Ni-Al Catalyst for Hydrogen Production via Methane Decomposition: Effect of Nickel Concentration and Spinel-like Structures. Int. J. Hydrog. Energy 2019, 44, 14424–14433. [Google Scholar] [CrossRef]
  25. Znak, L.; Zieliński, J. Effects of Support on Hydrogen Adsorption/Desorption on Nickel. Appl. Catal. A Gen. 2008, 334, 268–276. [Google Scholar] [CrossRef]
  26. Wang, T.; Tang, R.; Li, Z. Enhanced CO2 Methanation Activity over Ni/CeO2 Catalyst by Adjusting Metal-Support Interactions. Mol. Catal. 2024, 558, 114034. [Google Scholar] [CrossRef]
  27. Hao, Z.; Shen, J.; Lin, S.; Han, X.; Chang, X.; Liu, J.; Li, M.; Ma, X. Decoupling the Effect of Ni Particle Size and Surface Oxygen Deficiencies in CO2 Methanation over Ceria Supported Ni. Appl. Catal. B 2021, 286, 119922. [Google Scholar] [CrossRef]
  28. Szabados, M.; Szabados, T.; Mucsi, R.; Baán, K.; Kiss, J.; Szamosvölgyi, Á.; Sápi, A.; Kónya, Z.; Kukovecz, Á.; Sipos, P. Directed Thermocatalytic CO2 Reduction over NiAl4 Layered Double Hydroxide Precursors − Activity and Selectivity Control Using Different Interlayer Anions. J. CO2 Util. 2023, 75, 102567. [Google Scholar] [CrossRef]
  29. Ren, J.; Lei, H.; Mebrahtu, C.; Zeng, F.; Zheng, X.; Pei, G.; Zhang, W.; Wang, Z. Ni-Based Hydrotalcite-Derived Catalysts for Enhanced CO2 Methanation: Thermal Tuning of the Metal-Support Interaction. Appl. Catal. B 2024, 340, 123245. [Google Scholar] [CrossRef]
  30. Dębek, R.; Motak, M.; Duraczyska, D.; Launay, F.; Galvez, M.E.; Grzybek, T.; Da Costa, P. Methane Dry Reforming over Hydrotalcite-Derived Ni–Mg–Al Mixed Oxides: The Influence of Ni Content on Catalytic Activity, Selectivity and Stability. Catal. Sci. Technol. 2016, 6, 6705–6715. [Google Scholar] [CrossRef]
  31. Aziz, M.A.A.; Jalil, A.A.; Wongsakulphasatch, S.; Vo, D.-V.N. Understanding the Role of Surface Basic Sites of Catalysts in CO2 Activation in Dry Reforming of Methane: A Short Review. Catal. Sci. Technol. 2020, 10, 35–45. [Google Scholar] [CrossRef]
  32. Lima, D.d.S.; Dias, Y.R.; Perez-Lopez, O.W. CO2 Methanation over Ni–Al and Co–Al LDH-Derived Catalysts: The Role of Basicity. Sustain. Energy Fuels 2020, 4, 5747–5756. [Google Scholar] [CrossRef]
  33. Cao, M.; Li, S.; Wang, S.; Xu, W.; Zhou, X.; Ma, G.; Wang, X.; Nie, L.; Chen, Y. Designing Highly Active Hydrotalcite-Derived NiAl Catalysts for Methane Cracking to H2. Fuel 2024, 375, 132606. [Google Scholar] [CrossRef]
  34. Kaydouh, M.-N.; El Hassan, N.; Osman, A.I.; Ahmed, H.; Alarifi, N.; Fakeeha, A.H.; Bin Jumah, A.; Al-Fatesh, A.S. Optimizing CO2 Methanation: Effect of Surface Basicity and Active Phase Reducibility on Ni-Based Catalysts. React. Chem. Eng. 2024, 9, 1933–1946. [Google Scholar] [CrossRef]
  35. Xu, Y.; Du, X.; Shi, L.; Chen, T.; Wan, H.; Wang, P.; Wei, S.; Yao, B.; Zhu, J.; Song, M. Improved Performance of Ni/Al2O3 Catalyst Deriving from the Hydrotalcite Precursor Synthesized on Al2O3 Support for Dry Reforming of Methane. Int. J. Hydrog. Energy 2021, 46, 14301–14310. [Google Scholar] [CrossRef]
  36. Muniz, F.T.L.; Miranda, M.A.R.; Morilla dos Santos, C.; Sasaki, J.M. The Scherrer Equation and the Dynamical Theory of X-Ray Diffraction. Acta Crystallogr. A Found. Adv. 2016, 72, 385–390. [Google Scholar] [CrossRef]
  37. Gouveia, L.G.T.; Agustini, C.B.; Perez-Lopez, O.W.; Gutterres, M. CO2 Adsorption Using Solids with Different Surface and Acid-Base Properties. J. Environ. Chem. Eng. 2020, 8, 103823. [Google Scholar] [CrossRef]
  38. Stangeland, K.; Kalai, D.Y.; Li, H.; Yu, Z. Active and Stable Ni Based Catalysts and Processes for Biogas Upgrading: The Effect of Temperature and Initial Methane Concentration on CO2 Methanation. Appl. Energy 2018, 227, 206–212. [Google Scholar] [CrossRef]
Figure 1. Physicochemical properties in N2 physisorption of NiAl catalyst calcined at different temperatures: (a) Ni70Al30 400 °C for 8 h; (b) Ni70Al30 600 °C for 6 h; and (c) Ni70Al30 800 °C for 2h.
Figure 1. Physicochemical properties in N2 physisorption of NiAl catalyst calcined at different temperatures: (a) Ni70Al30 400 °C for 8 h; (b) Ni70Al30 600 °C for 6 h; and (c) Ni70Al30 800 °C for 2h.
Catalysts 15 00760 g001
Figure 2. (a) XRD patterns of calcined catalysts at different temperatures; (b) XRD patterns of sample calcined in different temperatures and reduced at 600 °C over 1h with H2.
Figure 2. (a) XRD patterns of calcined catalysts at different temperatures; (b) XRD patterns of sample calcined in different temperatures and reduced at 600 °C over 1h with H2.
Catalysts 15 00760 g002
Figure 3. H2-TPR profiles of calcined Ni-Al catalysts: (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h; (c) NiAl at 800 °C for 2h.
Figure 3. H2-TPR profiles of calcined Ni-Al catalysts: (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h; (c) NiAl at 800 °C for 2h.
Catalysts 15 00760 g003
Figure 4. H2-TPD profiles of calcined Ni-Al catalysts (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h (c) NiAl at 800 °C for 2h.
Figure 4. H2-TPD profiles of calcined Ni-Al catalysts (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h (c) NiAl at 800 °C for 2h.
Catalysts 15 00760 g004
Figure 5. CO2-TPD profiles of calcined Ni-Al catalyst. (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h; (c) NiAl at 800 °C for 2h.
Figure 5. CO2-TPD profiles of calcined Ni-Al catalyst. (a) NiAl at 400 °C for 8h; (b) NiAl at 600 °C for 6h; (c) NiAl at 800 °C for 2h.
Catalysts 15 00760 g005
Figure 6. Reaction results (a) CO2 conversion as a function of reaction temperature, sources of equilibrium curve and (b) XRD patterns and average crystallite size of spent catalysts.
Figure 6. Reaction results (a) CO2 conversion as a function of reaction temperature, sources of equilibrium curve and (b) XRD patterns and average crystallite size of spent catalysts.
Catalysts 15 00760 g006
Figure 7. (a) TPO and (b) DTA profiles of the spent catalysts after the methanation tests.
Figure 7. (a) TPO and (b) DTA profiles of the spent catalysts after the methanation tests.
Catalysts 15 00760 g007
Figure 8. Reaction results for (a) CO2 conversion under different reduction conditions and (b) TPO of spent catalysts and (c) XRD patterns of spent catalysts under different activation conditions.
Figure 8. Reaction results for (a) CO2 conversion under different reduction conditions and (b) TPO of spent catalysts and (c) XRD patterns of spent catalysts under different activation conditions.
Catalysts 15 00760 g008
Table 1. Physicochemical properties by N2 physisorption of Ni-Al catalyst calcined at different temperatures, and average crystallite size.
Table 1. Physicochemical properties by N2 physisorption of Ni-Al catalyst calcined at different temperatures, and average crystallite size.
SampleSBET (m2 g−1)Vpore (cm3 g−1)Dpore (nm)Crystallite Size (nm)
Calcined *Reduced **
NiAl 400 °C 8 h3080.3819.143.04.4
NiAl 600 °C 6 h2780.364.793.76.7
NiAl 800 °C 2 h2340.346.044.47.2
* Determined for NiO from reflection at 43.2° (200). ** Determined for Ni0 from reflection at 44.5° (111).
Table 2. Total dispersion, metallic area, and H2 consumption for reduced catalysts.
Table 2. Total dispersion, metallic area, and H2 consumption for reduced catalysts.
CatalystRelative Area (%) *Total H2
Consumption (µmol/gcat) *
Metallic Area (m2/g) **Total
Dispersion (%) **
Peak 1Peak 2Peak 3Peak 4
NiAl 400 °C 8 h4.948.612.534.0178.622.614.06
NiAl 600 °C 6 h0.05.274.120.796.523.414.21
NiAl 800 °C 2 h10.810.431.047.8177.812.512.25
* Obtained from H2-TPR. ** Obtained from H2-TPD.
Table 3. Total basicity and deconvolution of CO2-TPD profiles for reduced catalysts.
Table 3. Total basicity and deconvolution of CO2-TPD profiles for reduced catalysts.
SamplesTotal Basicity [μmol/g]Weak [μmol/g] *Medium [μmol/g] *Strong [μmol/g] *Very Strong [μmol/g] *
NiAl 400 °C 8 h 53.9111.88 (22)10.93 (20)16.02 (30)15.08 (28)
NiAl 600 °C 6 h44.4111.62 (26)10.31 (23)18.74 (42)3.75 (8)
NiAl 800 °C 2 h55.8411.24 (20)14.05 (25)19.30 (35)11.24 (20)
* The relative fraction of the peak in the deconvolution is shown in parentheses.
Table 4. Average crystallite size of spent catalysts under different activation conditions.
Table 4. Average crystallite size of spent catalysts under different activation conditions.
Temperature (°C)Time (h)Atmosphere ReductionCrystallite Size (nm)
400490 N2 + 10 H23.1
500290 N2 + 10 H23.9
600190 N2 + 10 H23.4
600190 N2 + 10 CH49.8
600190 N2 + 10 CO23.2
600175 N2 + 15 CO2 + 10 CH413.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Martins, N.; Perez-Lopez, O.W. Influence of Calcination and Reduction Conditions of Ni-Al-LDH Catalysts for CO2 Methanation. Catalysts 2025, 15, 760. https://doi.org/10.3390/catal15080760

AMA Style

Martins N, Perez-Lopez OW. Influence of Calcination and Reduction Conditions of Ni-Al-LDH Catalysts for CO2 Methanation. Catalysts. 2025; 15(8):760. https://doi.org/10.3390/catal15080760

Chicago/Turabian Style

Martins, Nailma, and Oscar W. Perez-Lopez. 2025. "Influence of Calcination and Reduction Conditions of Ni-Al-LDH Catalysts for CO2 Methanation" Catalysts 15, no. 8: 760. https://doi.org/10.3390/catal15080760

APA Style

Martins, N., & Perez-Lopez, O. W. (2025). Influence of Calcination and Reduction Conditions of Ni-Al-LDH Catalysts for CO2 Methanation. Catalysts, 15(8), 760. https://doi.org/10.3390/catal15080760

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop